You are on page 1of 4

Journal of Magnetism and Magnetic Materials 344 (2013) 207–210

Contents lists available at SciVerse ScienceDirect

Journal of Magnetism and Magnetic Materials


journal homepage: www.elsevier.com/locate/jmmm

Adiabatic temperature change around coinciding first and second


order magnetic transitions in Mn3Ga(C0.85N0.15)
Ö. Çakır a, M. Acet b,n
a
Physics Department, Yıldız Technical University, Istanbul, Turkey
b
Experimentalphysik, Universität Duisburg-Essen, D-47048 Duisburg, Germany

art ic l e i nf o a b s t r a c t

Article history: The inverse magnetocaloric material in Mn3GaC relies on the presence of a first order antiferromagnetic–
Received 18 February 2013 ferromagnetic transition at around 160 K. Since Mn3GaN is antiferromagnetic, the partial substitution of
Received in revised form carbon by nitrogen in Mn3GaC enhances the antiferromagnetic exchange and shifts this transition to
28 May 2013
higher temperatures. At the Mn3Ga(C0.85N0.15) stoichiometry, the transition takes place at 180 K. The
Available online 11 June 2013
hysteresis at the transition reduces so that the inverse magnetocaloric effect is practically reversible. We
Keywords: study the magnetocaloric effect and the reversibility of the adiabatic temperature-change by magnetiza-
Magnetocaloric effect tion and direct temperature-change measurements up to field-changes of 5 T.
Magnetic entropy & 2013 Elsevier B.V. All rights reserved.
Adiabatic temperature change
Antiperovskite

Magnetic refrigeration technology based on the magnetocaloric thermal range of the magnetostructural transition for magnetic
effect (MCE) has attracted much attention because of its potential refrigerant technology applications [7].
advantages such as environment-friendly and energy-efficient Among the Mn-based antiperovskites, Mn3GaC has been parti-
virtues compared to traditional refrigeration technology. Gener- cularly investigated in detail both from fundamental physical and
ally, materials to be used in magnetic refrigeration technology technological points of view. It undergoes a first-order antiferro-
should demonstrate large magnetocaloric effects [1]. Large con- magnetic (AF) to ferromagnetic (FM) transition at Tt ¼165 K and a
ventional or inverse magnetocaloric effects associated with large second-order FM to paramagnetic (PM) transition at TC ¼250 K
entropy changes ΔS are usually observed around first-order [12–18]. Without a change in the structure, an anomalous unit-cell
magnetostructural transitions [2–5]. These are larger than those volume-increase by about 0.5% occurs when the temperature is
observed around second-order magnetic transitions. However, the lowered through the AF–FM transition. Much effort has been spent
disadvantage for practical applications of magnetic refrigeration to shift the temperatures to chosen refrigeration temperature-
materials relying on magnetostructural transitions is the transition ranges. For example, partial replacement of Mn by Fe or Ni on
hysteresis which causes irreversibilities in the adiabatic substitutional sites shifts Tt to lower temperatures while shifting
temperature-change ΔT ad due to hysteresis losses. Nevertheless, TC to higher temperatures [19,20]. In Mn3AlCx varying the carbon
if the thermal hysteresis is sufficiently narrow, reversibility in ΔT ad content also varies Tt and TC [21]. The partial substitution of Ga by
can be established [6,7]. Al introduces new magnetic states for which the transition
Mn-based antiperovskite compounds in the form Mn3M C and temperatures can be shifted by an external magnetic field [22].
Mn3M N (M: Al, Zn, Ga, Ge, Sn, In) show a large variety of magnetic Pressure can also shift the transition temperatures [23]. The
ordering configurations and structural transitions and have been studies mentioned above have focused on the substitution of Mn
extensively studied [8–11]. These materials are ordered in a cubic by another 3d element or the substitution of Ga by another main
symmetry with the Pm3m space group, whereby the C or N atoms group element. Studies on the partial replacement of the carbon
are located at the octahedral site at the cube-center. In some cases, metalloid interstitial with nitrogen or boron are more limited in
the hysteresis related to the transition in these materials can also Mn-based antiperovskite materials [24–26].
be small, so that it is particularly worth investigating various Mn3GaN is antiferromagnetic (AF) with an isotropic structure
compositional combinations of these materials to tailor the [27], and earlier studies have shown that it has been possible to
control both Tt and TC of Mn3GaC by replacing the carbon
interstitial with nitrogen [24]. It has been shown that at a
concentration around x ¼0.10–0.20 in Mn3Ga(C1−xNx), Tt and TC
n
Corresponding author. approach each other and practically coincide so that on increasing
E-mail address: mehmet.acet@uni-due.de.(M. Acet) temperature, the transition takes place from an AF state to an

0304-8853/$ - see front matter & 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jmmm.2013.05.057
208 Ö. Çakır, M. Acet / Journal of Magnetism and Magnetic Materials 344 (2013) 207–210

enhanced paramagnetic (EPM) state, without any long-range FM


ordering setting in. The absence of long-range FM ordering could
lead to a decrease in the transition hysteresis and lower transi-
tional losses. On the other hand, a high-enough external magnetic-
field could stabilize long-range FM ordering, so that a field-
induced AF–FM transition can involve large entropy-changes
giving rise to large inverse magnetocaloric effects. Therefore, we
focus here on the hysteretic and magnetocaloric properties of a
sample with Mn3Ga(C0.85N0.15) stoichiometry and focus on the
reversibility of the direct adiabatic temperature-change, ΔT ad .
Direct ΔT ad measurements and the reversibility of ΔT ad on
repeatedly applying and removing the external magnetic field
give the ultimate test on the usefulness of the materials for MCE
applications. We investigate the magnetocaloric effect by deter-
mining the entropy-change from isothermal magnetization mea-
surements, M(H), and by direct adiabatic temperature-change,
Fig. 2. Temperature dependence of the magnetization under 5 mT (right scale) and
ΔT ad , measurements. ΔT ad was measured in a calorimeter with under 1, 3, and 5 T (left scale). Tt shifts with the applied field at a rate of −2.5 K T−1.
standard calorimetric techniques employing waiting times of The right-hand scale is for the data taken under 0.005 T.
about 10 min before and after applying the magnetic field to
account for nonadiabatic conditions. The magnetization measure-
ments were carried out using a superconducting quantum inter-
ference device magnetometer, and ΔT ad was measured directly in
a purpose-built calorimeter incorporating a 5 T superconducting
magnet.
First, Mn2N was obtained by flowing nitrogen gas over Mn
powder for 2 days at 850 1C. This was then mixed with Mn and C
powder and Ga pieces. They were pressed into a pellet and sealed
in a quartz tube under argon atmosphere. The sample was held at
800 1C for 5 days. X-ray powder diffraction (Fig. 1) verified the
antiperovskite structure with a lattice constant of 3.892 Å which is
somewhat smaller than the lattice constant 3.894 Å of Mn3GaC
[12,24]. A reflection related to excess graphite is indicated in the
figure.
The temperature-dependence of the magnetization M(T) is
shown in Fig. 2 in the temperature-range 5 ≤T ≤ 380 K and in
external magnetic fields of 0.005, 1, 3, and 5 T. The measurements
were taken with standard zero-field-cooled (ZFC) and field-cooled
(FC) protocols. As the temperature decreases from 380 K, M
increases tending to approach a FM state. However, before TC is
reached a transition to an AF state occurs at Tt, below which the
magnetization rapidly decreases. Tt shifts to lower temperatures
by about −2.5 K T−1. Since Tt and TC nearly coincide there is no
temperature-range for which long-range ferromagnetism is stable.

Fig. 3. The magnetization isotherms. (a) The isotherms in the vicinity of TI. (b) The
isotherm at 175 K obtained on increasing and decreasing field showing very weak
hysteresis, a property favorable for magnetic refrigeration.

The transition occurs between an AF and an EPM state. The data


taken under the higher fields show some residual ferromagnetism
so that M(T) is off-set close to 10 Am2 kg−1, unchanging under
measurement fields of 1–5 T, indicating that this value is about the
saturation value of the residual ferromagnet (this can also be seen
in Fig. 3 where M(H) initially rises steeply and eventually tends
toward saturation). Residual ferromagnetism acts as a conven-
tional magnetocaloric material, and therefore as a heat source
when a magnetic field is applied. This affects the measured ΔT ad of
the principal inverse magnetocaloric material in the sample, in a
manner that the temperature decrease on applying the field
becomes lower than the expected full value which a sample
Fig. 1. The X-ray diffraction pattern of Mn3Ga(C0.85N0.15). The reflections are related
without residual ferromagnetism would show. In Fig. 2, other
to the cubic antiperovskite structure. A reflection related to excess graphite is than the small difference in M(T) between the FC and ZFC states
indicated in the figure. for 0.005 T, no splitting is found in the data obtained in higher
Ö. Çakır, M. Acet / Journal of Magnetism and Magnetic Materials 344 (2013) 207–210 209

Fig. 4. The temperature-dependence of the entropy-change calculated from


magnetization isotherms for field-changes of 1–4.8 T. Fig. 5. Temperature dependence of the adiabatic temperature-change in a field-
change of 3 T. Prior to each measurement, the sample was taken initially to 160 K.
The lines are guides.
fields. Thermal hysteresis is essentially absent so that this material
could be useful for magnetic refrigeration.
The magnetization isotherms for Mn3Ga(C0.85N0.15) are shown
in Fig. 3 in the temperature-range of the transition, where
metamagnetic-like behavior indicates the presence of field-
induced transformations from the AF state to the EPM state
(Fig. 3(a)). The initial rise in M(H) at low fields is due to residual
ferromagnetism. Its presence leads to a lower estimated value of
ΔS from the numerical integration of the Maxwell relationships. In
Fig. 3(b), we show M(H) on increasing and decreasing fields where
there is very little difference showing that the field-induced
transitional hysteresis is very weak. A similar weak hysteresis
would also be expected in the transition range if the magnetiza-
tion could be measured under adiabatic conditions.
The temperature-dependence of the entropy-change ΔS is
shown in Fig. 4 for field changes ranging from 1 to 4.8 T. ΔS is
positive in the range of the transition exhibiting the inverse
magnetocaloric effect and is negative above it showing the con-
ventional magnetocaloric effect. Around 4 T, ΔS reaches its max-
imum value of about 9 J kg−1 K−1.
To investigate the magnetocaloric effect more closely and to
examine the suitability of the sample as an inverse magnetocaloric
material we have measured the temperature-dependence of ΔT ad
directly in an adiabatic calorimeter. The temperature of the sample
is initially brought to 160 K and then raised to the desired
measuring temperature. The sample is then isolated from the
surroundings and the temperature is monitored for 600 s to
account for nonadiabatic conditions. The field is then increased
to 3 T in about 200 s, and the temperature is continuously
̃ K. ΔT ad
Fig. 6. Effect of field-cycling. (a) Field-cycling of a 5 T-magnetic field at 176
recorded. After the field reaches its maximum value, the tempera-
is reversible with a value of about 3 K. The lines through the data guide the
ture is further monitored for another 600 s, again to account for
determination of ΔT ad by standard calorimetric techniques. (b) Applying and
nonadiabatic conditions. Fig. 5 shows that at low temperatures, removing a field of 3 T. Up to 2.1 T the inverse magnetocaloric effect is present.
there is practically no temperature-change. In the region of the For μ0 H 42:1 T, the conventional magnetocaloric effect occurs.
transition, the inverse magnetocaloric effect is observed: the
sample cools on applying a field and warms on removing it. The
absolute value of ΔT is essentially the same when the field is magnetic refrigeration, however it is important to note here that
applied or removed at any given measuring temperature. This is the sample may not be useful at all temperatures within the
consistent with the narrow hysteresis related to the transition transition region. As shown in Fig. 6b, where the initial tempera-
seen in Figs. 2 and 3(b). At higher temperatures the sample is in ture is 182.1 K and where ΔT ad undergoes a 0–3–0 T cycle, the
the EPM state, and a reversible conventional magnetocaloric effect sample first shows the inverse magnetocaloric effect with a
of about 1 K is observed. temperature-drop of 0.43 K up to about 2.1 T, but then the sample
To examine the reversibility of ΔT ad closer, we have measured it warms again by 0.36 K as the field further increases to 3 T. On
cyclicly at about 176 K, which is close to the maximum value of removing the field at the 1500 s point, the process reverses. The
ΔT ad in Fig. 6. Fig. 6(a) shows the 0–5–0 T cyclic application of the cause of this effect is due to the shift of the transition temperature
magnetic field over several periods. The absolute value of ΔT ad due to applied field to a value lower than that of the measuring
remains constant at a value of about 3 K over all periods due to the temperature, so that the system enters into the EPM state already
narrow hysteresis in this sample. This property is favorable for at 2.1 T and shows the conventional magnetocaloric effect.
210 Ö. Çakır, M. Acet / Journal of Magnetism and Magnetic Materials 344 (2013) 207–210

Therefore the sample warms when the field is further increased up Acknowledgments
from 2.1 to 3 T.
The cause of adiabatic cooling on applying a magnetic field, i.e., We thank S. Kılıç-Çetin for experimental assistance. This work
the inverse magnetocaloric effect is not fully understood. In the was supported by Deutsche Forschungsgemeinschaft (No.
case of the conventional magnetocaloric effect, applying a mag- SPP1599).
netic field causes a decrease in the magnetic entropy which is
balanced by an increase in the vibrational entropy accounting for
the observed warming, provided that the electronic entropy does
References
not change. A similar argument accounting for adiabatic cooling on
applying a magnetic field is not as straight forward. Adiabatic
[1] A.M. Tishin, Y.I. Spichkin, The Magnetocaloric Effect and its Application, IOP
cooling would require a decrease in the vibrational entropy. This Publishing, Bristol, 2003.
would then require the magnetic entropy to increase—which is [2] V.K. Pecharsky, K.A. Gschneidner Jr., Physical Review Letters 78 (1997) 4494.
not possible other than in a paraprocess where the adiabatic [3] T. Krenke, E. Duman, M. Acet, E.F. Wassermann, X. Moya, L. Ma nosa, A. Planes,
Nature Materials 4 (2005) 450.
temperature-change is in the order of a tenth of a degree [1].
[4] A. Planes, L. Mañosa, M. Acet, Journal of Physics: Condensed Matter 21 (2009)
Therefore, instead of attempting to account for the observed 233201.
cooling, one can base the argument on the fact that cooling is [5] E. Brück, N.T. Trung, Z.Q. Ou, K.H.J. Buschow, Scripta Materialia 67 (2012) 590.
observed and provide an explanation for the possible changes in [6] V.V. Khovaylo, K.P. Skokov, O. Gutfleisch, H. Miki, R. Kainuma, T. Kanomata,
Applied Physics Letters 97 (2010) 052503.
the various contributions to the entropy. [7] Ö. Çakır, M. Acet, Applied Physics Letters 100 (2012) 202404.
In an adiabatic process the total entropy-change ΔStot is zero, [8] J.-P. Bouchaud, R. Fruchart, R. Pauthenet, M. Guillot, H. Bartholin, F. Chaissé,
i.e., Journal of Applied Physics 37 (1966) 971.
[9] D. Fruchart, E.F. Bertaut, Journal of the Physical Society of Japan 44 (1978) 781.
ΔStot ¼ ΔSvib þ ΔSm þ ΔSel ¼ 0; ð1Þ [10] Ph. L'Héritier, D. Fruchart, R. Madar, R. Fruchart, Materials Research Bulletin 14
(1979) 1089.
where ΔSvib is the change in the vibrational entropy, ΔSm is the [11] D. Fruchart, Ph. L'Héritier, R. Fruchart, Materials Research Bulletin 15 (1980)
change in the magnetic entropy, and ΔSel is the change in the 415.
electronic entropy. Since cooling is observed in the inverse [12] T. Tohei, H. Wada, T. Kanomata, Journal of Applied Physics 94 (2003) 1800.
[13] M.H. Yu, L.H. Lewis, A.R. Moodenbaugh, Journal of Applied Physics 93 (2003)
magnetocaloric effect, Svib should decrease so that ΔSvib o 0. This 10128.
means that the sum of the other two terms should be positive [14] R. Burriel, L. Tocado, E. Palacios, T. Tohei, H. Wada, Journal of Magnetism and
since the total entropy-change must be zero in an adiabatic Magnetic Materials 290–291 (2005) 715.
[15] Ming-hui Yu, L.H. Lewis, A.R. Moodenbaugh, Journal of Magnetism and
process. Indeed, if the system undergoes a thermally induced
Magnetic Materials 299 (2006) 317.
magnetic transition from an AF to an EPM state, it cannot be [16] B.S. Wang, P. Tong, Y.P. Sun, X. Luo, X.B. Zhu, G. Li, X.D. Zhu, S.B. Zhang,
expected that ΔSel vanishes [7], so that the sum of ΔSm and ΔSel Z.R. Yang, W.H. Song, J.M. Dai, Europhysics Letters 85 (2009) 47004.
[17] B.S. Wang, P. Tong, Y.P. Sun, X.B. Zhu, X. Luo, G.Li, W.H. Song, Z.R. Yang, J.M. Dai,
has to compensate the observed decrease in ΔSvib to satisfy the
Journal of Applied Physics 105 (2009) 083907.
condition that ΔStot ¼ 0. What the individual signs of ΔSm and ΔSel [18] B.S. Wang, C.C. Li, J.C. Lin, L.J. Li, P. Tong, X.B. Zhu, Z.R. Yang, W.H. Song, J.M. Dai,
are cannot be deduced from these experiments. This could be one Y.P. Sun, Journal of Magnetism and Magnetic Materials 323 (2011) 2017.
way of looking at the inverse magnetocaloric effect problem. [19] B.S. Wang, P. Tong, Y.P. Sun, W. Tang, L.J. Li, X.B. Zhu, Z.R. Yang, W.H. Song,
Physica B 405 (2010) 2427.
However, we acknowledge the complexity of the issue of under- [20] B.S. Wang, P. Tong, Y.P. Sun, W. Tang, L.J. Li, X.B. Zhu, Z.R. Yang, W.H. Song,
standing the inverse magnetocaloric effect around a magnetos- Journal of Magnetism and Magnetic Materials 322 (2010) 163.
tructural phase transition especially when it comes to dealing with [21] P. Tong, Y.P. Sun, B.C. Zhao, X.B. Zhu, W.H. Song, Solid State Communications 14
(1979) 1203.
the separability of the various entropy contributions.
[22] K. Koyama, T. Kanomata, T. Watanabe, T. Suzuki, H. Nishihara, K. Watanabe,
Replacing carbon by nitrogen in Mn3GaC can increase Tt and Materials Transactions 47 (2006) 492.
simultaneously decrease TC and narrow the transitional hysteresis. [23] T. Kanomata, T. Kawashima, H. Yoshida, T. Kaneko, Journal of Applied Physics
The narrowing of the transitional hysteresis affects favorably the 67 (1990) 4824.
[24] Ph. L'Héritier, D. Boursier, R. Fruchart, D. Fruchart, Materials Research Bulletin
reversibility of the MCE on cyclicly applying and removing 14 (1979) 1203.
external magnetic fields. It is our further aim to undertake studies [25] Hu Jing-Yu, Wen Yong-Chun, Yao Yuan, Wang Cong, Zhao Qing, Jin Chang-
involving the replacement of carbon with other interstitial metal- Qing, Yu Ri-Cheng, Chinese Physics Letters 29 (2012) 086201.
[26] Y. Sun, Y. Guo, Y. Tsujimoto, J. Yang, B. Shen, W. Yi, Y. Matsushita, C. Wang,
loids to enhance and tailor the transitional properties of Mn-based
X. Wang, J. Li, C.I. Sathish, K. Yamaura, Inorganic Chemistry 52 (2013) 800.
antiperovskite compounds, especially in relation to the narrowing [27] E.F. Bertaut, D. Fruchart, J.P. Bouchaud, R. Fruchart, Solid State Communica-
of the transitional hysteresis. tions 6 (1968) 251.

You might also like